κ-carbide in a high-Mn light-weight steel: precipitation, off...

148
κ-carbide in a high-Mn light-weight steel: precipitation, off-stoichiometry and deformation Von der Fakultät für Georessourcen und Materialtechnik der Rheinisch-Westfälischen Technischen Hochschule Aachen zur Erlangung des akademischen Grades einer Doktorin der Ingenieurwissenschaften genehmigte Dissertation vorgelegt von M.Sc. Mengji Yao aus Zhuji, VR China Berichter: Prof. Dr.-Ing. Dierk Raabe Univ.-Prof. Jochen M. Schneider, Ph.D. Tag der mündlichen Prüfung: 20. März 2017 Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar

Transcript of κ-carbide in a high-Mn light-weight steel: precipitation, off...

  • κ-carbide in a high-Mn light-weight steel:

    precipitation, off-stoichiometry and deformation

    Von der Fakultät für Georessourcen und Materialtechnik

    der Rheinisch-Westfälischen Technischen Hochschule Aachen

    zur Erlangung des akademischen Grades einer

    Doktorin der Ingenieurwissenschaften

    genehmigte Dissertation

    vorgelegt von M.Sc.

    Mengji Yao

    aus Zhuji, VR China

    Berichter: Prof. Dr.-Ing. Dierk Raabe

    Univ.-Prof. Jochen M. Schneider, Ph.D.

    Tag der mündlichen Prüfung: 20. März 2017

    Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar

  • Acknowledgements

    The completion of this dissertation would not be possible without the people

    mentioned in the following. I would like to express my sincere gratitude

    To Prof. Dierk Raabe for offering me this project and providing guidance and

    encouragement throughout this work.

    To Prof. Jochen M. Schneider for his interest in this work and being my second

    supervisor.

    To Dr. Pyuck-Pa Choi, Dr. Michael Herbig and Dr. Baptiste Gault for the intensive

    discussions and valuable instructions.

    To Dr. Emanuel Welsch, Dr. Marta Lipinska-Chwalek, Dr. Christian Liebscher and

    Prof. Christina Scheu for their contributions to the transmission electron microscopy

    work.

    To Dr. Poulumi Dey, Dr. Tilmann Hickel and Prof. Jörg Neugebauer for their help

    on ab-initio calculations.

    To Dr. Philippe T. Pinard and Dr. Wenwen Song for performing electron probe

    micro-analysis and synchrotron x-ray diffraction experiments, respectively.

    To Mr. Andreas Sturm and Mr. Uwe Tezins for their support on the focused ion

    beam and atom probe tomography.

    To Dr. Hauke Springer, Mr. Frank Rütters and Mr. Frank Schlüter for the material

    synthesis and processing.

    To Ms. Monika Nellessen, Ms. Katja Agenendet and Ms. Heidi Bögershausen for

    their support to the metallography lab and scanning electron microscopy devices.

  • To all colleagues from Max-Planck-Institut für Eisenforschung GmbH, especially

    in the group of “Atom Probe Tomography” and in the office room 688, for the lovely

    working atmosphere and daily interactions.

    To my family and friends for their understanding and support all along this path.

    To the funding from European Research Council under the EU’s 7th Framework

    Programme (FP7/2007-2013)/ERC Grant agreement 290998 “SmartMet”.

  • Contents

    Symbols and abbreviations .......................................................................................... i

    1 Introduction ............................................................................................................ 1

    2 Materials and methods .......................................................................................... 6

    2.1 Material processing ........................................................................................... 6

    2.2 Microstructure characterization ......................................................................... 7

    2.2.1 Synchrotron X-ray diffraction .................................................................. 7

    2.2.2 Scanning electron microscopy ................................................................. 8

    2.2.3 Transmission electron microscopy .......................................................... 9

    2.2.4 Atom probe tomography .......................................................................... 9

    2.2.5 Correlative transmission electron microscopy and atom probe

    tomography .......................................................................................................... 12

    2.3 Density functional theory study ...................................................................... 15

    3 Dislocation-particle interaction during plastic deformation ............................ 17

    3.1 Introduction ..................................................................................................... 17

    3.2 Results .............................................................................................................. 19

    3.2.1 Morphology and arrangement of κ-carbides .......................................... 19

  • 3.2.2 Dislocation/κ-carbide interaction ........................................................... 21

    3.3 Discussion ...................................................................................................... 26

    3.3.1 Particle shearing and ordering strengthening ......................................... 26

    3.3.2 Particle dissolution and solute segregation ............................................ 28

    4 Elemental partitioning and κ/ interface structure ........................................... 31

    4.1 Introduction ................................................................................................... 31

    4.2 Results ........................................................................................................... 32

    4.2.1 κ/ elemental partitioning and phase compositions ................................ 32

    4.2.2 κ/ interface structure ............................................................................. 37

    4.3 Discussion ...................................................................................................... 44

    4.3.1 Compositional accuracy of APT measurements .................................... 44

    5 Off-stoichiometry and site-occupancy of κ-carbides ......................................... 60

    5.1 Introduction ................................................................................................... 60

    5.2 Results ........................................................................................................... 61

    5.2.1 APT-based analysis of site-occupancy ................................................... 61

    5.2.2 DFT-based investigations of off-stoichiometry and site-occupancy ...... 71

    5.3 Discussion ...................................................................................................... 76

    5.3.1 Off-stoichiometry and site-occupancy of κ-carbides ............................. 76

    5.3.2 Underlying issues related to DFT ........................................................... 78

    6 Phase equilibria in the κ/ microstructure ......................................................... 80

    6.1 Introduction ................................................................................................... 80

    6.2 Results ........................................................................................................... 81

    6.2.1 Prolonged-aged microstructure .............................................................. 81

    6.2.2 Coarsening of grain interior precipitates ................................................ 85

    6.2.3 Grain boundary discontinuous precipitation .......................................... 89

  • 6.3 Discussion ..................................................................................................... 97

    6.3.1 Influence of strain state on the off-stoichiometry of κ-carbides ............ 97

    6.3.2 Thermal stability of coherent GI κ-carbide precipitates ........................ 98

    6.3.3 Grain boundary discontinuous precipitation .......................................... 99

    7 Summary and outlook ....................................................................................... 107

    7.1 Summary ..................................................................................................... 107

    7.2 Outlook ........................................................................................................ 109

    Bibliography ............................................................................................................. 110

    List of tables .............................................................................................................. 121

    List of figures ............................................................................................................ 123

    Curriculum Vitae ..................................................................................................... 131

    Abstract ..................................................................................................................... 133

    Zusammenfassung .................................................................................................... 135

  • i

    Symbols and abbreviations

    Symbols

    Ferrite

    Lattice misfit

    True strain

    pPartitioning coefficient

    Austenite

    𝛾𝐴𝑃𝐵

    Anti-phase boundary energy

    ’ Gamma-prime phase in superalloys

    κ Kappa-carbide

    μ Chemical potential

    Diffraction angle

    𝜎 Stress

    r Particle radius

    a Lattice parameter

    b Burgers vector

    c Composition

    d{hkl} Interplanar distance of the {hkl} lattice planes

    D Diffusion coefficient

    E Formation energy

  • ii

    f Mole fraction

    G Shear modulus

    kB Boltzmann constant

    M Taylor factor

    t Time

    T Temperature

    Vf Volume fraction

    X Diffusion length

    Z Atomic number

    Abbreviations

    1D One-dimensional

    2D Two-dimensional

    3D Three-dimensional

    APB Anti-phase boundary

    APT Atom probe tomography

    BSE Backscattered electron

    DF Dark field

    DP Discontinuous precipitation

    EBSD Electron backscatter diffraction

    ECCI Electron channeling contrast imaging

    EDX Electron dispersive x-ray spectroscopy

    EPMA Electron probe microanalysis

    ER Evaporation rate

    fcc Face-centered cubic

    FEG Field emission gun

    FM Ferromagnetic

    FIB Focused ion beam

    FOV Field-of-view

    GB Grain boundary

    GI Grain interior

    HAADF High angle annular dark field

  • iii

    HAGB High angle grain boundary

    HR High-resolution

    LEAP Local electrode atom probe

    LRO Long-range ordering

    MBIP Microband-induced plasticity

    OM Optical microscopy

    OPS Oxide polishing suspension

    PF Pulse fraction

    RDF Radial distribution function

    ROI Region of interest

    SAD Selected area diffraction

    SDM Spatial distribution map

    SE Secondary electron

    SEM Scanning electron microscopy

    SFE Stacking fault energy

    SIP Shear-band induced plasticity

    SQS Quasi random structure

    SRO Short-range ordering

    SS Solid-solution

    STEM Scanning transmission electron microscopy

    SXRD Synchrotron X-ray diffraction

    TEM Transmission electron microscopy

    TOF Time-of-flight

    TRIP Transformation induced plasticity

    TWIP Twinning induced plasticity

    XRD X-ray diffraction

    ZA Zone axis

  • iv

  • 1

    Chapter 1

    Introduction

    There is a strong demand for the development of advanced high-strength steels for

    automotive applications, in order to reduce energy consumption and greenhouse gas

    emission. Austenitic () high-Mn (15-30 wt.%) Fe-Mn-Al-C steels show particularly

    outstanding mechanical properties (Frommeyer et al., 2006; Chang et al., 2010;

    Springer et al., 2012; Gutierrez-Urrutia et al., 2013; Park et al., 2013; Raabe et al., 2014;

    Rana et al., 2014) and are therefore highly promising candidates for such applications.

    Due to their good oxidation and corrosion resistance (Banerji et al., 1978), these steels

    were originally developed in an attempt to substitute Cr-containing stainless steels, but

    recently they have regained substantial interest due to their excellent strength-ductility

    balance and significantly reduced mass density because of alloying with Al.

    Austenitic Fe-Mn-Al-C steels, containing high Al (5-12 wt.%) and high C (0.5-1.3

    wt.%), are age-hardenable. Through an ageing treatment at a temperature of 470-710 °C

    (Sato et al., 1989; Tjong et al., 1990; Sato et al., 1990; Choo et al., 1997), nanometer-

    scale ordered κ-carbide precipitates can be introduced into the disordered face-centered

    cubic (fcc) austenite matrix, which can remarkably strengthen the material while

    preserving great ductility. Due to the great importance of κ-carbide precipitates,

    numerous studies have been conducted to understand their precipitation and

    deformation behavior, which, however, are not fully clear.

  • Chapter 1. Introduction

    2

    Pronounced planar slip is generally observed in deformed κ-carbide-containing -

    alloys (Gutierrez-Urrutia et al., 2013; Park et al., 2013; Raabe et al., 2014; Welsch,

    2016). Though a few transmission electron microscopy (TEM) micrographs of sheared

    particles have been reported (Choi et al., 2010; Welsch, 2016), controversial opinions

    still exist over the predominant interaction mechanism between κ-carbide precipitates

    and dislocations. Gutierrez-Urrutia et al. concluded that the primary mechanisms are

    Orowan bypassing of particle stacks and subsequent expansion of dislocation loops

    assisted by cross-slip (Gutierrez-Urrutia et al., 2014). Since the κ-carbide precipitates

    are finely dispersed in the -matrix, it is challenging to make a clear observation of them

    in the presence of dislocations. The current existing two-dimensional (2D) TEM

    micrographs are not convincing enough.

    Moreover, no matter which mechanism dominates, a better microstructural

    characterization is yet necessary to quantitatively analyze the strengthening behavior.

    As for the “Orowan looping” case, the particle size and inter-spacing are critical which

    however are not unambiguous in 2D projected micrographs. In the case of “particle

    shearing”, besides particle size, its feasibility is associated with the antiphase boundary

    (APB) energy on the slip plane (Ardell et al., 1988), which depends on the stoichiometry

    and site-occupancy of the ordered particle.

    Commonly, the composition of κ-carbides is given as (Fe,Mn)3AlCx, where the

    exact chemical composition of this phase is still unknown (Sato et al., 1988; Tjong,

    1990; Sato et al., 1990; Choo et al., 1997). It is tacitly assumed to be a derivative from

    the Fe3AlCx-type ternary κ-carbide (Oshima et al., 1972). Due to the difficulty in

    determining C concentrations in small precipitates and also obtaining single-phase κ-

    carbide, the exact composition of the Fe3AlCx-type κ-carbide was uncertain for a long

    time (Palm et al., 1995; Sanders et al., 1997). In 1995, Palm et al. revealed a

    composition range for this phase between Fe3.2Al0.8C0.71 and Fe2.8Al1.2C0.42, i.e.

    Fe3+yAl1-yCx (-0.2

  • Chapter 1. Introduction

    3

    few nanometers in size, their direct chemical characterization is challenging. Atom

    probe tomography (APT) is the ideal tool to resolve this composition uncertainty since

    it offers near-atomic spatial resolution and equal detection sensitivity to all elements

    (Miller et al., 1996; Gault et al., 2012b; Larson et al., 2013; Miller et al., 2014).

    However, in contrast to intensive studies on κ-carbide-containing alloys, little APT data

    have been published on κ-carbide precipitates in quaternary austenitic Fe-Mn-Al-C

    alloys.

    The crystal structure of κ-carbides is reported to be L’12 perovskite-type (Tjong,

    1986; Tjong, 1990; Han et al., 1986; Choo et al., 1985). In such a variant of the

    conventional fcc crystal structure, the unit cell contains 5 atoms at 3 kinds of sites. In

    addition to the 4 atoms at one corner and three face-centered sites, as per fcc

    conventional unit cell, there is also one atom at the body-centered site. Conventionally,

    as shown in Figure 1.1, the elemental site-occupancy of the κ-carbide phase is such that

    Al occupies the corner site of the unit cell, Fe and Mn the face-centered positions, and

    C occupies the body-centered octahedral interstitial site, which is based on observations

    of electron diffraction patterns and calculations of structure factors (Tjong, 1986; Han

    et al., 1986; Tjong, 1990). In terms of atomic percentage, the stoichiometric

    (Fe,Mn)3AlC κ-carbide has 60% (Fe+Mn), 20% Al and remaining 20% C. However,

    the κ-carbides are commonly expected to be off-stoichiometric and thus the elemental

    site-occupancy remains unresolved.

    Figure 1.1. Schematic sketch of the unit cell of the ideal stoichiometric (Fe,Mn)3AlC

    L’12 κ-carbide with Fe, Mn, Al and C atoms shown by red, orange, green and purple

    balls, respectively.

  • Chapter 1. Introduction

    4

    In addition, nanometer-sized κ-carbide precipitates are often regarded as coherent

    particles with respect to the -matrix according to x-ray diffraction (XRD) analyses

    (Huang et al., 1994) or simply tacitly assumed. The interfacial structure between κ-

    carbide precipitate and -matrix is little known, which, however, could significantly

    affect the morphology and arrangement of κ-carbide precipitates due to elastic constrain

    if any and the dislocation motion if misfit dislocation occurs.

    Besides the beneficial intragranular precipitation of κ-carbide, detrimental

    intergranular precipitation of coarse κ0-carbide via discontinuous precipitation (DP) has

    also been noticed in Fe-Mn-Al-C alloys (Chao et al., 1993; Hwang et al., 1993; Huang

    et al., 1994; Choo et al., 1997; Kimura et al., 2004). The resultant lamellae colonies,

    composed of κ0-carbide and other solute-depleted phase(s), gradually wipe out the GI

    structure by GB migration. Though the DP product phases were identified in these

    works, the initiation and evolution of such a phase transition has not been investigated.

    From the mechanical performance perspective, it is of great importance to understand

    this process. Also, from a thermodynamic point of view, the phase equilibria and

    elemental partitioning in such a microstructure need a better understanding.

    To address the questions discussed above, a quaternary model Fe-29.8Mn-7.7Al-

    1.3C (wt.%) steel was chosen in this work. An ageing treatment at 600 °C was utilized

    for the κ-carbide precipitation. The well-precipitated 24 hours-aged sample was

    subjected to tensile deformation in order to study the deformation mechanisms. Given

    that the nanometer-sized κ-carbide precipitates are dense and chemical analysis is

    compulsory for the phase composition and elemental partitioning study, three-

    dimensional (3D) APT is the optimal tool for the κ-carbide characterization, thus

    employed as the primary technique. Besides, synchrotron x-ray diffraction (SXRD),

    high-resolution scanning electron microscopy (SEM) and (scanning) transmission

    electron microscopy ((S)TEM) are also utilized to obtain a general overview of the

    microstructure or complementary information to APT results. Theoretical calculations

    in the framework of density functional theory (DFT) is additionally employed to

    understand the atomistic structure of the off-stoichiometric κ-carbide.

    This thesis is organized in the following way.

    Chapter 1: Introduction.

  • Chapter 1. Introduction

    5

    Chapter 2: Materials and methods. The details of material synthesis and thermos-

    mechanical processing is introduced in this chapter. The devices and methods

    utilized in this work for microstructure characterization and the theoretical

    approach are described in detail.

    Chapter 3: Dislocation-particle interaction during plastic deformation. In this chapter,

    the 3D morphology and arrangement of κ-carbides in -matrix are firstly studied by

    APT and the κ/ alloy deformed to different strains are studied via correlative TEM

    and APT technique, revealing the dislocation-particle interaction mechanism.

    Chapter 4: Elemental partitioning and κ/ interface structure. In this chapter, the κ/

    elemental partitioning and chemical compositions of κ/ phases are studied by APT,

    the compositional accuracy of which is discussed. The κ/ interfacial structure, in

    terms of chemical gradient, interfacial width and coherency, are investigated using

    SXRD, STEM and APT.

    Chapter 5: Off-stoichiometry and site-occupancy of κ-carbide. In this chapter, attempts

    are made to resolve the site-occupancy of off-stoichiometric κ-carbide from both

    the experimental and theoretical perspective. DFT calculations successfully predict

    the off-stoichiometry and explains such a phenomenon.

    Chapter 6: Phase equilibria in the κ/ microstructure. In this chapter, the prolonged-

    aged κ/ microstructure is characterized via SXRD, SEM and APT, revealing the

    co-existence of coherency equilibrium of GI κ/ phases and thermodynamic

    equilibrium of GB κ0/0/α-ferrite phases. The thermal stability of GI κ-carbide and

    initiation & evolution mechanisms of GB κ0/0/α-ferrite phases are discussed.

    Chapter 7: Summary and outlook.

  • 6

    Chapter 2

    Materials and methods

    2.1 Material processing

    A high-Mn Fe-30Mn-8Al-1.3C (wt.%) steel was synthesized as a rectangular bar

    (12 kg, thickness of 40 mm) via melting and casting in an induction furnace under argon

    atmosphere. Due to evaporation of the melt and its interaction with the crucible during

    the production process, extra amounts of the alloying elements were added in order to

    achieve a target composition. According to empirical experience, a Fe-30.8Mn-8.2Al-

    1.34C alloy was produced aiming at the final Fe-30.0Mn-8.0Al-1.3C composition (all

    in wt.%). The composition of the as-cast alloy, as determined by wet chemical analysis,

    is listed in Table 0.1.

    Table 0.1. Chemical composition of the studied alloy obtained by wet chemical analysis.

    Fe Mn Al C

    wt.% 61.2 29.8 7.74 1.28

    at. % 53.9 26.7 14.12 5.25

    In order to remove microstructural heterogeneities inherited from the solidification

    process, the cast ingot was reheated to 1200 °C for 30 min and subsequently hot-rolled

    at 1100 °C to a thickness reduction of 75%, followed by water-quenching. The bar

    sample of 11×11×60 mm3, cut from the hot-rolled sheet, was then subjected to a solid-

  • Chapter 2. Materials and methods

    7

    solution (SS) treatment at 1200 °C for 2 h in a furnace under argon atmosphere and

    quenched to room temperature in an oil bath. At 1200 °C, the alloy is fully austenitic

    (Chin et al., 2010).

    To induce the precipitation of κ-carbides and study their evolution and phase

    stabilities upon ageing, the SS-treated sample was isothermally annealed at 600 °C for

    various times from 24 hours up to three months and oil quenched. The aged samples

    are referred to ‘x-aged’, where ‘x’ indicates the ageing time.

    The 24h-aged alloy with well-precipitated nanometer-scale κ-carbides was chosen

    as a representative case for the study on the dislocation-particle interaction during

    plastic deformation. Cylindrical tensile test samples with a gauge dimension of Φ 6

    mm×40 mm were prepared and interruptedly tested to true strains of ε=0.02, 0.05 and

    0.15 in a Zwick ZH 100 tensile machine with an initial strain rate of 5×10-4 s-1.

    2.2 Microstructure characterization

    Microstructure characterization of various samples was conducted using multi-

    scale techniques, including synchrotron X-ray diffraction (SXRD), optical microscopy

    (OM), scanning electron microscopy (SEM), transmission electron microscopy (TEM)

    and atom probe tomography (APT).

    2.2.1 Synchrotron X-ray diffraction

    Due to the high coherency between the κ-carbide precipitates and -austenite

    matrix (Sato et al., 1988; Sato et al., 1989; Sato et al., 1990; Tsay et al., 2011), i.e. little

    lattice misfit, as well as potential grain boundary (GB) phases formed in small amounts

    (Chao et al., 1993; Hwang et al., 1993), it is difficult to identify phases and trace their

    evolution upon ageing by laboratory X-ray diffraction. High-resolution SXRD with a

    superior signal-to-background ratio was hence employed. Thin cylindrical film samples

    of 3 mm in diameter were machined from alloys under different ageing states and

    mechanically grinded by SiC abrasive papers of grits from 120 up to 2000 (European

    P-grade) till reaching a sample thickness of ~1 mm. The SXRD measurements were

    carried out by Wenwen Song from the Lehrstuhl und Institut für Eisenhüttenkunde

    (IEHK), Rheinisch-Westfälische Technische Hochschule Aachen (RWTH Aachen) at

  • Chapter 2. Materials and methods

    8

    the high-resolution powder diffraction beamline P02.1 PETRA III Bessy with a

    wavelength of 0.20727 Å. The detected 2D diffraction patterns were first calibrated and

    integrated into one-dimensional (1D) data (intensity versus 2θ) by the software Fit2D

    (Hammersley et al., 1994) and further analyzed by the Rietveld method using the

    software MAUD (Lutterotti et al., 1999).

    2.2.2 Scanning electron microscopy

    Samples for microscopy studies were all prepared following the standard

    metallographic procedure:

    firstly, grinded by SiC abrasive papers of grit from 120 up to 4000 (European P-

    grade) under flowing water;

    then mechanically polished using 3 μm-grained diamond suspension with

    lubricant;

    finally, polished by colloidal silica oxide polishing suspension (OPS) with a few

    drops of soap to remove mechanically induced surface deformation and ensure a

    clean surface.

    For the sake of simplicity during sample preparation and handling, disc samples

    machined from gauges of tensile test samples were embedded into thermosetting

    bakelite with carbon filler of standard cylindrical geometry (Φ 25 mm) by a hot-

    mounting process under 180 °C and 25 kN (Struers PolyFast). To better reveal the

    multi-phase microstructure, some samples were etched with a 1% Nital solution (a

    mixture of nitric acid and ethanol) for 30-90 seconds. Prior to electron microscopy

    analysis, OM was often used to get a rough overview of the microstructure.

    Further microstructural characterization was performed in two field emission gun

    (FEG) HR-SEMs, i.e. JEOL JSM 6500F (JEOL GmbH) and ZEISS Crossbeam XB

    1540 (Carl Zeiss SMT AG), both operated at 15 kV and equipped with EDAX electron

    backscatter diffraction (EBSD) system. The ZEISS microscope was mainly employed

    for secondary electron (SE) and backscattered electron (BSE) imaging, whereas most

    EBSD measurements were carried out in the JEOL tool. The EBSD data was acquired

    and analyzed using the TSL OIM software (version 6.5-7.0).

  • Chapter 2. Materials and methods

    9

    2.2.3 Transmission electron microscopy

    TEM was employed for nanometer-sized precipitate analysis. The TEM work,

    including phase analysis by selected area diffraction (SAD) and precipitate observation

    using dark-field (DF) imaging, was mainly completed by Emanuel Welsch from the

    Department of Microstructure Physics and Alloy Design, Max-Planck-Institut für

    Eisenforschung GmbH (MPIE) and reported in his PhD dissertation (Welsch, 2016).

    Marta Lipinska-Chwalek from the Peter Grünberg Institut (PGI), Forschungszentrum

    Jülich also kindly helped for Scanning-TEM (STEM) measurements on the coherency

    of κ-carbide precipitate with respect to the -matrix in a FEI Titan G2 80-200

    ChemiSTEM operated at 200 kV.

    High angle annular dark field (HAADF) STEM samples were prepared by focused

    ion beam (FIB) using two FIB/SEM dual beam devices - FEI Helios Nano-Lab 600 and

    FEI Helios Nano-Lab 600i. Since the κ-carbides have a {001} cube-cube orientation

    relationship with the -matrix (Choo et al., 1997), is the optimal direction for an

    edge-on view of the κ/ interface, which was therefore adopted as the primary analysis

    axis. Moreover, for a fcc-metal, dislocation slip is typically activated along

    directions on {111} planes, whose edge-on observation direction would be .

    Hence, thin lamellas for misfit dislocation study were prepared along {001} and {011}

    planes, i.e. lifted out along {001} and {011} plane traces in a {100}-oriented grain

    (plane normal), respectively. The aim of orientation-specific preparation is to simplify

    TEM operation, i.e. minimizing the tilting required to achieve target zone axes and the

    increase in sample thickness caused by tilting. The crystallographic directions of bulk

    samples for lift-out were determined by prior EBSD measurements. The FIB-prepared

    lamellae were first cleaned by oxygen plasma for 30 seconds before performing STEM

    studies.

    2.2.4 Atom probe tomography

    TEM/STEM is good at crystallography, phase identification, defect and strain

    analysis down to atomic scale whereas it is difficult to map the distribution of light

    elements like carbon with electron dispersive x-ray spectroscopy (EDX). TEM images

    are also 2D projections of a 3D microstructure, which could also cause an ambiguous

  • Chapter 2. Materials and methods

    10

    interpretation of precipitate morphologies and alignments. Alternatively, APT, a

    combination of ion projection microscopy and a time-of-flight (TOF) spectroscopy, has

    a near-atomic resolution with equal sensitivity to all elements (Gault et al., 2012b). It

    was thereby employed for a three-dimensional (3D) chemical mapping of the

    nanometer-sized κ-carbides embedded in the -matrix.

    Needle-shaped APT specimens were all prepared by FIB using the two FIB/SEM

    dual beam devices as mentioned above. The usual lift-out procedure (Thompson et al.,

    2007) was employed for most specimens. To minimize the Ga implantation into

    specimens, during specimen sharpening the ion beam energy was reduced from 30

    kV/~0.26 nA to 16 kV/~0.14 nA after the 1st annular ion milling step and 2 kV/ ~24 pA

    was adopted for the final cleaning.

    For GB studies, it is challenging to prepare a specimen with a GB in the analysis

    volume since GBs are probably inclined with respect to sample surface and not visible

    by FIB/SEM during annual ion milling. To increase the success rate of capturing a GB,

    the GB was aligned parallel to the analysis direction by a tilted lift-out and then simply

    kept as the center of the annular milling patterns. A pre-cutting deep into the bulk

    sample across GB was first made to obtain the GB inclination angle, i.e. the angle

    between the GB and the bulk sample’s surface. During the lift-out, the stage was then

    tilted by an angle complementary to the GB inclination angle, which directly made the

    GB perpendicular to the specimen post (parallel to the analysis direction).

    During the analysis of the reaction front of GB discontinuous precipitation (DP),

    considering the field-of-view (FOV) of an APT specimen is much larger longitudinally

    than laterally, the reaction front was oriented perpendicular to the analysis direction by

    the method described in (Felfer et al., 2012), to enlarge the analysis volume in front of

    and behind the reaction front. Commercial Mo-grids (1GM 100, Pyser-SGI) were used

    as specimen posts in this case to realize the 90° rotation, simply by switching the Mo-

    grid holder adaptor with two perpendicular stage-mounting pins. The preparation of

    Mo-grids and its corresponding special home-made holder and holder adapters for FIB

    and APT instruments are explained in detail in (Herbig et al., 2015a).

    Two local electrode atom probe (LEAP) devices were used in this work for APT

    measurements. The LEAP 3000X-HR (CAMECA instruments) with a reflectron has a

    detection efficiency of ~37% while the LEAP 5000X-S (CAMECA instruments), a

  • Chapter 2. Materials and methods

    11

    straight path system, has a high detection efficiency of ~80%. Atom probe instrument

    equipped with a reflectron is usually expected to have improved mass resolutions (the

    peak width in a mass spectrum normalized by the mass of the peak) (Cerezo et al.,

    1998). Most measurements were performed in the LEAP 3000X-HR tool. The LEAP

    5000X-S was used for a comparable study on the mass spectra and crystallography.

    Needle-shaped specimens were measured in voltage-pulsing mode at ~70 K with a

    pulse repetition of 200 kHz, a pulse fraction (PF) of 15% and a target evaporation rate

    (ER) of 5 ions per 1000 pulses. Choosing such a set of parameters is based on a detailed

    study of the influence of measurement parameters on phase compositions, which will

    be explained in detail in section 4.3.1.2. The collected APT data were reconstructed and

    analyzed using the IVAS software (version 3.6.6 - 3.6.14) by CAMECA instruments.

    Figure 2.1. APT mass spectrum of κ-carbide containing austenitic alloy obtained by

    LEAP 3000X-HR: (a) 5-15 Da, (b) 18-26 Da, (c) 26-30 Da, (d) 33-57 Da. Ranges of

    mass-to-charge-state ratios without peaks are omitted. Peaks with ion overlap are

    marked with constituent ions.

    Figure 2.1 shows a representative mass spectrum of the studied austenitic alloy

    with κ-carbide precipitation. Peaks at 6, 6.5, 12, 13 Da are identified as carbon

  • Chapter 2. Materials and methods

    12

    monomers. Apart from them, several peaks can be assigned to carbon molecular ions,

    similar to previous studies on other high-C containing alloys and carbides (Miyamoto

    et al., 2012; Marceau et al., 2013; Takahashi et al., 2011; Li et al., 2011; Kitaguchi et

    al., 2014; Thuvander et al., 2011). Peaks at 18, 18.5, 36 and 37 Da belong to carbon

    trimer ions. The carbon dimers 12C21+ and (13C12C)1+ are detected at 24 and 25 Da,

    respectively. The presence of 24.5 Da peak, which can be assigned to (13C12C3)2+,

    indicates the detection of carbon tetramers (Takahashi et al., 2011; Li et al., 2011; Sha

    et al., 1992). Hence, the peak at 24 Da does not only result from the carbon dimer 12C21+

    but also partly from 12C4 2+. This overlapped peak can be decomposed by considering

    the peak at 24.5 Da (13C12C3)2+. According to the natural abundance of carbon isotopes

    12C and 13C, the contribution of 12C42+ to peak 24 Da can be estimated by its relative

    abundance ratio to (13C12C3)2+ (Takahashi et al., 2011; Li et al., 2011; Sha et al., 1992).

    The decomposition via the IVAS software indicates that 12C42+ gives a contribution of

    about 37% to the peak at 24 Da. The 12C22+ peak could in principle overlap with the

    12C1+ peak at 12 Da. However, the absence of a peak belonging to (12C13C)2+ suggests

    that there is a negligible fraction of 12C22+ at the peak at 12 Da. Regarding substitutional

    elements, Fe2+ is detected at 27, 28, 28.5 and 29 Da and its primary isotope can also be

    detected at 56 Da (56Fe+) and 18.7 Da (56Fe3+). Mn2+ is detected at 27.5 Da. Peaks at 9

    and 13.5 Da can be ascribed to Al3+ and Al2+, respectively. Here, it is also noted that

    according to the natural abundance of the isotopes of Fe, the peak at 27 Da cannot be

    completely assigned to 54Fe2+. Al1+ ions must give a contribution to it as well (Seol et

    al., 2012; Seol et al., 2013). The peak decomposition algorithm in IVAS indicates that

    45% of this peak can be assigned to Al1+. The measured total impurity concentration Si

    (14 Da) and Cr (26 Da) is less than 0.1 at.%. The material is a rather clean quaternary

    Fe-Mn-Al-C model alloy.

    2.2.5 Correlative transmission electron microscopy and atom probe

    tomography

    The particle-matrix interface and deformed microstructure of the 24h-aged alloy

    were further studied by the correlative TEM/STEM and APT approach (Herbig et al.,

    2015a). Although APT is capable of mapping a 3D microstructure with sub-nm

    resolution, due to different evaporation fields of κ-carbides and -austenite, local

  • Chapter 2. Materials and methods

    13

    magnification could occur and complicate the 3D reconstruction of the κ/ interface

    (Gault et al., 2012b). Also, it is often difficult to resolve crystal defects, such as

    dislocations and slip bands, only by APT due to the lack of the crystallographic

    information. Combining TEM/STEM and APT enables a correlative and

    complementary understanding on the interface structure and particle-dislocation

    interaction.

    Electro-polished Mo-grids were used as specimen posts for the correlative study,

    which could be easily transferred between FIB, TEM and APT with a grid holder

    (Herbig et al., 2015a). Similar to STEM lamella analysis, to optimize the measurement

    conditions, needle-shaped specimens are preferred to be orientation-specific. Due to the

    {001} cube-cube κ/ orientation relationship, the specimens for κ/ interface study were

    prepared orthogonally along -directions and put into a double-tilting TEM holder

    in such a way that the (α, β) tilting axes and electron beam were all parallel to the

    direction. Thereby the electron beam would directly observe edge-on the {001} κ/

    interface. Also, there would be κ/ interfaces perpendicular to the APT analysis

    direction, minimizing the local magnification effect during field evaporation and

    utilizing the higher resolution along the in-depth direction (Gault et al., 2012b). Owing

    to the dense distribution of the κ-carbides, there are often enough precipitates for

    interface analysis within the FOV of STEM and APT, i.e. ~80 nm from apex, beyond

    which the specimen thickness would be too large for STEM analysis.

    In contrast, the correlative study on slip bands is challenging in terms of specimen

    preparation. Due to the limited FOV of APT and common premature failure of

    deformed APT specimen, the slip bands have to be within the top ~100 nm from the

    needle’s apex in order to be detected in APT before a specimen facture. This is

    challenging since there is no image contrast of slip bands at all during annular FIB

    milling. For heavily deformed samples, e.g. ε=0.15, it is possible to realize that, owing

    to the high density of slip bands with an average spacing of ~250 nm, as well as the

    activation of multiple groups of slip bands (Welsch, 2016). However, the slip bands in

    slightly deformed samples (ε

  • Chapter 2. Materials and methods

    14

    exposure during the zone axes search process affecting the specimen survival rate in

    subsequent APT measurements, an orientation-specific specimen is desirable.

    The optimal direction for observing particle shearing along typical {111} fcc slip

    planes is . The strategies employed for such a specimen preparation are as

    follows.

    Prior EBSD and electron channeling contrast imaging (ECCI) (Zaefferer et al.,

    2014) were utilized to search for one set of {011} dense sharp slip lines in a {011}-

    oriented grain (done by Emanuel Welsch). Slip lines parallel to the longest {011}

    plane trace were preferred since they are less inclined from the sample surface, i.e.

    with a higher probability of being captured in the analysis volume.

    The relative distance and angle between the slip lines and microstructural features

    such as GB and inclusions, which could be visible by FIB/SEM, were measured.

    The target set of slip lines was located by FIB/SEM according to its relative

    position with respect to the microstructural features.

    The lift-out was carried out in such a way that the long-axis of the wedge was

    perpendicular to the slip lines.

    Specimens prepared in this way would directly fit the TEM zone axis (ZA)

    and should include several inclined slip bands in the top volume. The subsequent

    TEM characterization could manifest the distance between slip bands and needle

    apex and a further controlled FIB re-sharpening could bring the target slip band to

    the FOV of APT measurement.

    In order to minimize beam contamination and surface oxidation, all the specimens

    were freshly prepared just before loading into TEM double-tilting holders. Christian

    Liebscher from Department of Structure and Nano-/ Micromechanics of Materials,

    MPIE and Marta Lipinska-Chwalek performed STEM measurements on κ/ interface

    in a probe-corrected FEI Titan 80-300 S/TEM microscopy operated at an accelerating

    voltage of 300 kV. The TEM observation on deformed specimens was done by Emanuel

    Welsch in a Philips CM20 device. After the TEM work, the specimens were directly

    transferred into the FIB for re-sharpening when necessary and a re-cleaning at 2 kV for

    a few seconds to remove the contamination and oxidized layer before APT

    measurements.

  • Chapter 2. Materials and methods

    15

    2.3 Density functional theory study

    To understand the site-occupancy of ordered κ-carbide and the influence of elastic

    strain on its stoichiometry, ab-initio work was employed in this work. Poulumi Dey and

    Tilmann Hickel from Department of computational materials design, MPIE performed

    the calculations using density functional theory (DFT) as implemented in the Vienna

    Ab Initio Simulation Package (VASP) (Kresse et al., 1993). Projector augmented wave

    (PAW) potentials were used to describe the electron-ion interaction (Blöchl, 1994). The

    generalized-gradient approximation (GGA) functional of Perdew, Burke and Ernzerhof

    (PBE) (Perdew et al., 1996) was employed. The single-electron wave functions were

    expanded using plane waves up to an energy cut-off of 500 eV. The Methfessel-Paxton

    method (Methfessel et al., 1989) was used for the Fermi surface smearing with a 6×6×6

    Monkhorst-Pack grid for 2×2×2 atomic supercells of both stoichiometric and off-

    stoichiometric κ-carbides. The energies converged to a precision of 1 meV/atom.

    Structural relaxations were performed until the forces on each atom were below 0.01

    eV/Å. Cell shape and atomic positions were fully relaxed in all calculations, unless

    specified otherwise.

    The 2×2×2 supercell of stoichiometric (Fe,Mn)3AlC κ-carbide is composed of 8

    unit cells in the L’12 structure with 5 sites/unit cell and occupied by 40 atoms

    ((Fe16Mn8)Al8C8). Here, Fe/Mn are situated at face-centered sites of each unit cell, Al

    at corner sites and C atoms occupy the body-centered octahedral sites (Figure 2.2(a)).

    An Fe to Mn ratio of 2:1 was employed based on the APT results. DFT studies reveal

    that configurational entropy is more favorable than chemical ordering in the face-

    centered Fe/Mn sub-lattice of κ-carbide above ~75 K (Dey et al., 2016). The Fe/Mn

    chemical disorder was taken into account by a special quasi-random structure (SQS)

    generation scheme (Zunger et al., 1990) for the 2×2×2 supercell. It is important to note

    that all calculations were performed for ferromagnetic (FM) κ-carbide since the (Mn)

    anti-site formation energies computed for FM and paramagnetic states differed only by

    a small amount (~0.05 eV). Therefore, the consideration of paramagnetism, which is

    computationally much more demanding, does not qualitatively alter the results on

    structural properties and estimated point defect concentrations in κ-carbide (Dey et al.,

    2016).

  • Chapter 2. Materials and methods

    16

    Figure 2.2. Schematic visualization of the supercell of (a) stoichiometric L’12 κ-carbide,

    (Fe16Mn8)Al8C8, and (b) off-stoichiometric κ-carbide with a Mn anti-site at the Al sub-

    lattice and C vacancies, (Fe16Mn8)(Mn1Al7)(C5Vac3), with chemical disorder in the

    Fe/Mn sub-lattice with an Fe:Mn ratio of 2:1. Fe, Mn, Al and C atoms are shown by

    red, orange, green and purple balls, respectively. The theoretical calculations were done

    by Poulumi Dey and Tilmann Hickel.

  • 17

    Chapter 3

    Dislocation-particle interaction

    during plastic deformation

    3.1 Introduction

    A common characteristic of alloys showing excellent strength and ductility is a

    pronounced strain hardening capability, which continuously increases the strength and

    delays local necking during deformation. Transformation induced plasticity (TRIP) and

    twinning induced plasticity (TWIP) steels are typical examples for alloys showing high

    strain hardening capabilities (Herrera et al., 2011; Gutierrez-Urrutia et al., 2011), where

    martensitic transformation and formation of deformation twins are the respective

    dominant deformation and strain hardening mechanisms. The active deformation

    mechanisms are closely related to the stacking fault energy (SFE). Generally, as the

    SFE increases, the dominant deformation mechanism changes from TRIP to TWIP and

    from TWIP to dislocation gliding (Allain et al., 2004; Pierce et al., 2014).

    Gutierrez-Urrutia et al. attributed the excellent strain hardening capacity of solid

    solution austenitic Fe-Mn-Al-C alloys containing 5 wt.% Al, intragranular precipitation of nanometer-sized κ-carbides leads to a

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    18

    substantial increase in yield strength without a significant loss in ductility. Planar

    dislocation substructures were observed in deformed κ-carbide containing austenitic Fe-

    M-Al-C alloys. The occurrence of planar dislocation substructures in these alloys,

    which have high SFE values ranging from 80 to 110 mJ/m2 (Park et al., 2013), was

    ascribed to shearing of ordered κ-carbides by dislocations and a mechanism referred to

    as “glide plane softening” (Gerold et al., 1989). According to this mechanism, the local

    order is destroyed by a leading dislocation shearing the ordered particle and thus

    facilitating the glide of consecutive trailing dislocations. For high strain levels, the

    mechanisms of shear-band-induced plasticity (SIP) (Frommeyer et al., 2006) and

    microband-induced plasticity (MBIP) (Yoo et al., 2009; Park et al., 2013) have been

    suggested, referring to homogeneous shear deformation and bands of very high

    dislocation density, respectively.

    However, the interaction between dislocations and κ-carbide precipitates in age-

    hardened austenitic Fe-Mn-Al-C steels is still not well understood. Debates still exist

    over the dominant co-deformation mechanism during plastic deformation, i.e. the

    competition between dislocation looping and precipitate shearing. Choi et al. claimed

    that κ-carbide precipitates are sheared by slip bands since misalignment of particle

    stacks was observed along a specific direction by TEM DF imaging (Choi et al., 2010).

    However, Gutierrez-Urrutia et al. suggested that Orowan bypassing of stacks of κ-

    carbides is predominant, followed by expansion of dislocation loops by cross-slip while

    shearing of κ-carbides occurs only rarely (Gutierrez-Urrutia et al., 2014). Convincing

    observations are still missing to illustrate the dislocation interaction with κ-carbide

    precipitates. The answer to this question is the prerequisite to understand the nature of

    the strengthening by κ-carbide precipitates and thereby accurately predict the yield

    strengths of the precipitation-hardened κ/ alloys.

    This chapter aims at providing a comprehensive understanding of the interaction

    of dislocations with κ-carbide precipitates, as well as the associated strengthening and

    deformation mechanisms. First of all, the 3D morphology and arrangement of κ-carbide

    precipitates in the non-deformed -matrix are unveiled by APT analysis. Then

    correlative TEM-APT analyses are employed to systematically characterize the

    deformed κ/ microstructure under various strain states. Crystallographic and chemical

    mapping are correlated to each other, demonstrating the dislocation interaction with κ-

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    19

    carbide precipitates. Finally, the strengthening effect of κ-carbide precipitates, how they

    are affected by plastic deformation and its reverse effects on deformation are discussed.

    The 24h-aged sample is chosen for this work as a state, for which precipitation is

    pronounced and is similar to the previously reported, typical κ/ microstructure (Choo

    et al., 1997; Choi et al., 2010; Sato et al., 1988-1990; Tjong et al., 1990).

    3.2 Results

    3.2.1 Morphology and arrangement of κ-carbides

    Precipitation of coherent κ-carbides in the -matrix introduces an elastic stress

    field due to the lattice mismatch between precipitates and matrix (Choo et al., 1997).

    The shape and arrangement of κ-precipitates are determined by the minimization of the

    sum of the elastic energy and the interfacial energy, arising from the κ/ phase

    boundaries (Doi et al., 1996).

    Figure 3.1(a) shows the ordered L’12-type κ-carbide precipitates within the

    austenitic matrix observed by APT. The κ-carbides in the reconstructed 3D atom maps

    are visualized by 9 at.% C iso-concentration surfaces. APT reveals that cuboidal and

    plate-like κ-carbide particles are arranged in 3D in the form of stacks along three

    orthogonal directions which are identified by TEM observations as the elastically soft

    crystallographic directions (Gutierrez-Urrutia et al., 2012; Gutierrez-Urrutia et

    al., 2014; Choo et al., 1997). The cuboidal particles have a size of 15-20 nm. The largest

    cross-sections of plate-like particles perpendicular to the particle stack directions are

    often square-shaped with an edge length of 15-20 nm, while the thickness of the plates

    is on the order of 5-10 nm. The spacing between two parallel stacks is on the order of

    10-40 nm and will be referred to as broad -channels. The spacing between individual

    precipitates within a stack is about 2-5 nm and will be referred to as narrow -channels.

    Figure 3.1(b) shows a schematic sketch of -carbide precipitates’ 3D morphology

    and arrangement based on APT observations. Two possible corresponding 2D transect-

    projections along zone axes, which are often employed for imaging -carbides

    with TEM, are shown in Figure 3.1(c). Particles that might be assumed to have a

    cuboidal morphology based on their square-like appearance in the TEM DF images

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    20

    (Gutierrez-Urrutia et al., 2012; Gutierrez-Urrutia et al., 2014; Choo et al., 1997) have

    very often a much smaller size in the third dimension according to APT. Thus, they

    show plate-like morphology (see the particles outlined in red). At the intersection

    between two stacks of precipitates a preference for a cuboidal morphology is often

    observed, e.g. the particles outlined in green. Long rectangular particles observed in

    TEM micrographs (particles highlighted in yellow outlines) are found to be rectangular

    parallelepipeds.

    Figure 3.1. Morphology and arrangement of κ-carbide precipitates in 3D space (APT)

    and 2D observation (e.g. TEM): (a) three representative reconstructed 3D APT maps of

    C (purple), Al (green) and Mn (yellow) atoms. The κ-carbide precipitates are visualized

    with a 9 at.% C iso-concentration surface. (b) A schematic illustration of the 3D

    morphology and arrangement of κ-carbide precipitates based on APT observations. (c)

    2D projections of the κ/ microstructure along directions highlighted in (b),

    simulating the TEM observation.

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    21

    3.2.2 Dislocation/κ-carbide interaction

    Figure 3.2. TEM dark-field (DF) images of deformed microstructures at different true

    strains of (a) ε=0.05 and (b) ε=0.15 utilizing the superlattice diffraction spot of κ-

    carbides. Two different zone axes (ZA) (a) [011] and (b) [001] were employed to show

    the shearing of κ-carbides and sheared κ-carbides, respectively. The TEM

    measurements were carried out by Emanuel Welsch.

    The deformed microstructure of the 24h-aged samples are shown in Figure 3.2,

    which displays TEM micrographs obtained from APT specimens. The TEM

    observations were obtained by Emanuel Welsch. At a low strain level of 0.05, clear

    shearing of κ-carbide stacks is observed along the [011] ZA as indicated by arrows in

    Figure 3.2(a). The κ-carbide precipitates have a (001) cube-cube orientation relationship

    with the -matrix. Along the [011] ZA, i.e. looking edge-on the (11 1̅) plane, the

    nanometer-sized κ-carbides with a 2-5 nm inter-spacing in the narrow -channels

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    22

    overlap with each other in the 2D TEM projection. Instead of individual precipitates, a

    stack of precipitates is observed. The misalignments of several stacks along a single slip

    line is captured (Figure 3.2(a)), clearly revealing the shearing of κ-carbides by slip. As

    the true strain increases to 0.15, more slip systems are activated and intensive slip bands

    appear (highlighted by arrows in Figure 3.2(b)), which, along the [001] ZA, are

    manifested by dense precipitate fragments concentrating on certain directions. In

    comparison to the relatively complete precipitates at the apex of the needle-shaped

    specimen, the κ-carbide precipitates within these slip bands are fragmented into small

    debris. At the intersection of slip bands, the superlattice phase contrast is almost

    completely lost, indicating either the loss of ordering or even dissolution of precipitates.

    It is worth noting that from low-strain to high-strain states, slip bands are observed

    throughout the specimens, not only in these two representative ones shown in the Figure

    3.2, but also in another approx. 10 specimens. Particle shearing is a general phenomenon,

    observed in the deformed κ/ microstructure, and therefore believed to be the prevalent

    deformation mechanism. The previous debates in the literature on either shearing or

    Orowan looping is probably due to the lack of clear observation of such deformed

    microstructures. The high density of particles and specific crystallographic orientation

    add difficulties onto that. Here, these difficulties are circumvented by optimized

    specimen preparation, adopting thin needle-shaped specimens specifically oriented with

    the help of a prior EBSD measurement.

    The deformed microstructure with 3D chemical mapping, accomplished by

    correlative TEM -APT on a 0.05-strained specimen, is presented in Figure 3.3. In the

    TEM micrograph (Figure 3.3(a)), two originally cuboidal or plate-like κ-carbide

    precipitates are clearly split into two parts - one major part and another smaller one at

    the corner, as highlighted by blue arrows. The slight mismatch between these two parts,

    particularly for the right plate-like κ-carbide, indicates that these two precipitates are

    sheared by dislocation slip. The slip directions of the two precipitates are parallel to

    each other. Since the needle-shaped specimen is prepared orthogonally along the

    direction, the slip direction can be characterized as ~45° from [001] and [010], i.e. close

    to [011]. Besides, as highlighted by yellow arrows, there are also some small fragments

    of the ordered precipitates, which probably have been heavily fragmented by

    deformation.

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    23

    Figure 3.3. Correlative TEM and APT analysis of a deformed microstructure at a true

    strain of ε=0.15: (a) TEM dark-field (DF) micrograph along [001] zone axis with the

    blue arrows highlighting the shearing of κ-carbides and yellow ones indicating the

    fragments of κ-carbides; (b) an overlay of reconstructed APT analysis volume with

    purple κ-carbides visualized by carbon iso-concentration surface at a threshold value of

    9.0 at.% on top of the TEM DF micrograph of the same specimen taken before APT

    measurement; (c-e) 3D elemental atom maps of the reconstructed APT volume with

    different elemental iso-concentration surfaces of (c) C 9.0 at.%, (d) C 7.5 at.% and

    (e) Al 14.5 at.%, showing the dissolution of κ-carbides. The TEM measurements were

    carried out by Emanuel Welsch.

    The same specimen was subsequently analyzed by APT. Figure 3.3(b) shows an

    overlay of the reconstructed APT analysis volume on top of the DF-TEM micrograph.

    The latter utilizes the superlattice diffraction contrast to reveal the ordered κ-carbide

    precipitates whereas the former employs carbon iso-concentration surfaces at a

    threshold value of 9.0 at.% to show the carbon-enriched κ-carbides. The size, shape and

    position of κ-carbide precipitates in the reconstructed APT volume fit well with the DF-

    TEM image, illustrating that the employed APT reconstruction algorithm, tip profile

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    24

    reconstruction, could perform properly for this κ/ alloy. Despite the well reconstructed

    APT volume, particle splitting by dislocation slip is not clearly observed, conversely to

    what was highlighted by blue arrows in the DF-TEM image (Figure 3.3(a)). This is

    probably due to shearing of the precipitate by a single dislocation glide rather than dense

    slip bands of high dislocation density, as captured in the TEM micrograph. The signal

    is likely smoothed out by voxelization and delocalization of APT data. Instead of two

    split parts of one precipitate, only one complete κ-carbide is visualized by carbon iso-

    concentration surface in the APT volume. In addition to large κ-carbides both observed

    in TEM and APT, there are also small, irregularly shaped fragments with no obvious

    visual match between TEM and APT measurements: the region of which is marked by

    a yellow dash circle in Figure 3.3(b). Some fragments observed in the DF-TEM image

    are not found in APT analysis via 9.0 at.% C iso-concentration surface, and vice versa.

    This is probably due to the limited FOV of APT so that the outer shell of the specimen

    observed in TEM are not registered at the APT detector (Herbig et al., 2015a). Or the

    κ-carbides at this region fragmented by deformation preserve local ordering and solute

    enrichment to some extent, but not necessarily both at the same time.

    To better reveal the 3D solute distribution after deformation, further APT

    elemental maps and iso-concentration surfaces of different elements at varying

    threshold values are created (Figure 3.3(c-e)). As highlighted by a black dashed circle

    in Figure 3.3(c), the elements are not homogeneously distributed at this region. A lower

    threshold value for the C iso-concentration surface of 7.5 at.% unveils that there are not

    only small C-enriched fragments but also C segregation along certain directions, as

    indicated by black arrows in Figure 3.3(d). Such a linear solute enrichment is also

    correspondingly found for Al at the same positions (Figure 3.3(e)), but not for Mn (not

    shown here). Completely differing from the common particle morphology (Figure 3.1),

    these line features imply that the κ-carbides in this region have been fragmented and

    dissolved during deformation. The dislocation slip might destroy the κ-carbides and

    drag the solutes along with its movement. Considering the affinity between segregated

    solutes and crystalline defects (Kirchheim et al., 2007), it is highly possible that the

    solute segregation zones are dislocation lines. The reason why such an enrichment only

    takes place for Al and C, not Mn, is that the κ-carbides are enriched with Al and C, not

    Mn (details for κ/ elemental partitioning are referred to Chapter 4.2.1).

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    25

    Owing to the -oriented specimen, as well as the (001) cube-cube orientation

    relationship between κ-carbide precipitates and -matrix, the directions can be

    easily figured out for the reconstructed 3D APT analysis volume (Figure 3.4). Therefore,

    a crystallographic analysis can be performed on the linear solute segregation zones,

    highlighted by black arrows in Figure 3.3(d).

    Figure 3.4. Crystallographic analysis of three solute-segregated line features in a

    deformed κ/ alloy at a true strain of 0.15 as highlighted by black arrows in Figure 3.3(d)

    based on the orientation of the needle-shaped specimen.

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    26

    All three line segregation zones (Figure 3.4(a), (b) and (c)) are well associated to

    {111} planes, which are the typical slip plane for fcc metals. Along with the

    aforementioned slip direction (Figure 3.3(a)), the activated slip system can be

    identified as {111}, that is the typical slip system for fcc metals. Among these

    three linear segregation zones, one line is found to be perpendicular to the slip

    direction (Figure 3.4(a)) and the other two are inclined to that (Figure 3.4(b)(c)), which

    implies that probably the former is an edge dislocation while the latter have a mixed

    character.

    3.3 Discussion

    3.3.1 Particle shearing and ordering strengthening

    The microstructure of the 24h-aged alloy deformed to different strain states have

    been studied by correlative TEM and APT. It is revealed that {111} dislocation

    slip and shearing of κ-carbide precipitates occurs during plastic deformation rather than

    Orowan looping.

    The ordered κ-carbide with a L’12-type perovskite structure deform differently

    from the disordered fcc -matrix. For dislocation slip on a crystallographic plane, a

    shear stress, the so-called Peierls stress, has to be exerted on the slip plane, which

    depends exponentially on the ratio of lattice spacing d to the Burgers vector b and

    thereby ought to be smallest for a slip along the close-packed direction on the close-

    packed plane (Gottstein, 2004). In disordered fcc , this corresponds to the

    a/2{111} slip system (a is the lattice parameter). However, in chemically ordered

    L’12 κ-carbides the preferred slip system could be a{001} or a{111}, since

    the shortest lattice vectors a do not lie in the close-packed {111} planes.

    Experimentally, the observed activated slip system is the latter one. Given that the

    Burgers vector of a perfect dislocation in the -matrix (a/2) is only half of the

    closing vector to restore the ordered κ-carbide to its perfect lattice, it cannot enter the

    κ-carbide unless a planar defect is formed (Ardell, 1985; Reed, 2008). A pair of

    superpartial a/2{111} dislocations must travel together through the κ-carbides,

    forming the so-called superdislocation. The resultant planar fault in-between two

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    27

    superpartials is known as an anti-phase boundary (APB) and the APB energy 𝛾𝐴𝑃𝐵

    represents the associated energy barrier for the occurrence of particle cutting.

    Depending on 𝛾𝐴𝑃𝐵, volume fraction Vf and the mean particle radius r (assumed

    spherical for simplicity) of the κ-carbides, the pair of superpartial dislocations could

    reside in the same κ-carbides or there could be faulted particles between the leading and

    trailing superpartials, i.e. ‘strongly coupled’ or ‘weakly coupled’ superpartials,

    respectively (Reed, 2008). Their anticipated order strengthening contribution are

    different considering the elastic stress between the superpartials. In the SS-treated alloy,

    free of precipitates, the spacing between the superpartials intersecting the short-range

    ordering (SRO) was found to be ~20 nm (Welsch et al., 2016). Since the APB energy

    is usually relatively high and should be much higher in the 24h-aged alloy with κ-

    carbide precipitates of long-range ordering (LRO) than that in SS-treated alloy, the

    superpartials would experience more resistance in the κ-carbide containing alloy. The

    superpartial spacing should be less than ~20 nm and comparable to the particle size.

    Hence, the superdislocation in this case is more inclined to strong-coupling, whose

    ordering strengthening contribution is

    ∆𝜎𝑜𝑟𝑑𝑒𝑟𝑖𝑛𝑔 = √3

    2 𝑀 (

    𝐺𝑏

    𝑟) 𝑉𝑓

    1

    2 𝑤

    𝜋32

    (2𝜋 𝛾𝐴𝑃𝐵 𝑟

    𝑤 𝐺 𝑏2− 1)

    1

    2, (3.1)

    where M is the Taylor factor of 3.06 to convert the resolved shear strength for

    dislocation motion into an equivalent uniaxial yield strength of a polycrystal (Stoller et

    al., 2000), G represents the shear modulus, b denotes the Burgers vector and w is a

    dimensionless constant of the order of unity. No 𝛾𝐴𝑃𝐵 data is reported specifically for

    κ-carbides, but similarities with superalloys with a L12 ordered structure and the

    activated {111} slip system (Reed, 2008)) allow us to use values reported for

    those alloys (𝛾𝐴𝑃𝐵 ~0.1 J/m2) as a rough estimate. Taking G=70 GPa, b=0.26 nm

    (Welsch et al., 2016), r=10 nm (half of the particle size ~ 20nm) and 𝑉𝑓 ~0.21

    (experimental results refer to section 4.2), ∆𝜎𝑜𝑟𝑑𝑒𝑟𝑖𝑛𝑔 is estimated to be 321 MPa. This

    is much lower and more reasonable than what is estimated for Orowan stress

    ∆𝜎𝑂𝑟𝑜𝑤𝑎𝑛~𝑀 𝐺 𝑏 √𝑉𝑓

    𝑟=2.55 GPa (Gottstein, 2004), supporting that particle cutting is the

    preferred mechanism rather than Orowan bypassing.

    Experimentally, the ageing-introduced κ-carbides was found to significantly

    increase the yielding strength of the alloy by ~480 MPa (Welsch, 2016). Given that the

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    28

    two samples, SS-treated and 24h-aged, have similar grain sizes and the solute

    concentration is lower in the aged alloy as a result of precipitation, the order

    strengthening contribution, estimated as 321 MPa for κ-carbide shearing, is relatively

    low in comparison to the strength increase of ~480 MPa. One reason could be that there

    are additional contributions arising from the elastic strain field due to coherency, shear

    modulus mismatch between κ-carbide precipitates and -matrix, and creation of

    additional matrix-precipitate interface during particle shearing (Ardell, 1985). On the

    other hand, a few assumptions are made for the estimation of order strengthening.

    Firstly, for the sake of simplicity, the κ-carbides are assumed to be homogeneously

    distributed. In fact, as shown in Figure 3.1, they form particle stacks with different

    widths for broad and narrow -channels. Locally, along the particle stacks, the volume

    fraction is much higher than the averaged ~0.21. Secondly, there is actually no reported

    data of 𝛾𝐴𝑃𝐵 for κ-carbides. ~0.1 J/m2 is only based on the ’ precipitates in superalloys

    with a similar crystal structure and a similar slip system, which may lead to an

    inaccurately estimated strength increase.

    3.3.2 Particle dissolution and solute segregation

    Despite having a high SFE, the 24h-aged alloy exhibits pronounced planar slip

    during deformation (Welsch, 2016). It has been ascribed to the so-called glide plane

    softening phenomenon, which states that the cutting of ordered precipitates on a single

    slip plane by superdislocations results in a continuous decrease of obstacle strength on

    this individual plane and only minor strength needs to be overcome by the following

    dislocation pairs until all particles of the activated slip plane are completely cut (Gerold

    et al., 1989). Thereby, dislocation motion and particle shearing is restricted to the slip

    plane, giving rise to the planar slip with isolated active slip planes. As strain increases,

    multiple slip bands are activated and the spacing between slip bands is continuously

    reduced (Welsch, 2016). The refinement of planar slip bands progressively destroys the

    κ-carbides (Figure 3.2-3.3). Inside the slip bands, particularly at their intersections, κ-

    carbides are found fragmented or even dissolved.

    Similar dissolution of shearable particles under plastic deformation has been

    reported in severe plastic deformed metals, e.g. dissolution of metastable fine

    precipitates in equal-channel angular pressed Al alloys (Murayama et al., 2001;

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    29

    Gutierrez-Urrutia et al., 2005), and in fatigued metals, e.g. precipitate dissolution within

    the persistent slip bands or shear bands (Vogel et al., 1982; Vinogradov et al., 2002).

    In those cases, the intensive strain, either applied by severe deformation or accumulated

    by cycling, cut the particles into smaller nanoscale fragments, leading to the instability

    of the phase because of high surface energy and their eventual dissolution. Such a

    particle fragmentation is also observed in the present κ/ alloy by massive slip bands in

    high-strain states (Figure 3.2-3.3), which promotes the dissolution of κ-carbides.

    It is known, on the other hand, that the dislocation interaction with solute atoms

    of precipitates might also cause the dissolution or decomposition of particles. One well-

    studied case is the cementite decomposition in heavily drawn pearlitic steels (Li et al.,

    2011). During the motion of dislocations, looping around the interfaces in that case,

    carbon atoms are dragged out of the cementite and segregate to dislocation structures.

    As for shearable particles, the penetration of dislocations through particles might lead

    to the transfer of interstitial atoms from the particles to the matrix due to drifting of

    atoms under the stress field of dislocations (Sagaradze et al., 1997). In the present case,

    the observed solute-segregated line features are likely dislocations, the crystallographic

    orientations of which follows the typical fcc slip system (Figure 3.4). A compositional

    analysis of such a segregation is illustrated in Figure 3.5 using proxigram (proximity

    histogram, Hellman et al., 2000) and 1D concentration profile. Figure 3.5(b) shows the

    proxigram generated around the line segregations highlighted in purple in Figure 3.5 (a)

    by a 7.5 at.% C iso-concentration surface. The 1D concentration profiles of cylindrical

    regions of interest (ROIs) placed perpendicular to dislocation lines, as shown in Figure

    3.5(a), are plotted in Figure 3.5(c-d). Both methods clearly reveal the enrichment of C

    and Al to the dislocation lines, the enrichment factor of which are about 2.8 and 1.2 in

    comparison to the neighborhood compositions, respectively. This result suggests that

    upon particle cutting, indeed the solutes of the κ-carbide precipitates are dragged along

    with the moving dislocations, which facilitates precipitate fragmentation and

    dissolution. Regarding the composition, no obvious difference is observed between the

    two solute-segregated dislocation lines in Figure 3.5(c) and Figure 3.5(d), which are

    probably of edge and mixed characters according to the analysis in Figure 3.4(a) and

    Figure 3.4(b), respectively.

  • Chapter 3. Dislocation-particle interaction during plastic defromation

    30

    Figure 3.5. Concentration analysis on solute segregation to dislocations: (b) Proxigram

    of the magenta-highlighted interface generated by a carbon iso-concentration surface of

    7.5 at.% in (a). (c)(d) Two individual 1D concentration profiles of cylindrical regions

    of interest (ROIs) perpendicular to two dislocation lines as highlighted in (a).

  • 31

    Chapter 4

    Elemental partitioning and κ/

    interface structure

    4.1 Introduction

    The precipitation of nanometer-sized κ-carbides in austenitic Fe-Mn-Al-C alloys

    significantly strengthens these grades while maintaining substantial formability. This is

    realized by a dramatic increase in yield strength and an excellent strain hardening ability

    comparable to that of κ-carbide-free alloys (Welsch, 2016). Since particle-shearing is

    found to be the dominating deformation mechanism, the significant strengthening effect

    results from ordering strengthening, as well as contributions of coherency, shear

    modulus mismatch and creation of additional matrix-precipitate interface (section 3.3.1).

    In order to better understand the strengthening and strain hardening behavior of the

    studied alloy, it is essential to analyze the internal κ/ microstructure, especially the

    elemental partitioning and the interfaces between the precipitates and the matrix.

    The partitioning of solutes is decisive for the lattice misfit between the κ-carbides

    and the -matrix. In a coherent system, elastic strain energy arises from this mismatch

    and the precipitates experience an elastic strain field. The critical influence of the elastic

    strain field on resultant morphology and arrangement of coherent precipitates has been

    noticed in many alloy systems (Doi, 1996; Lee, 1996). As the lattice misfit increases,

  • Chapter 4. Elemental partitioning and κ/ interface structure

    32

    the shape of the particles may evolve from sphere to cube and plate, or even needle. The

    corresponding elastic anisotropy can lead to a directional alignment of particles instead

    of a random distribution. In this chapter, APT is employed for the chemical mapping of

    the κ/ microstructure at the atomic scale, illustrating the elemental partitioning between

    phases and measuring the chemical composition of κ-carbides. The initial state before

    ageing (SS-treated) and the 24h-aged alloy with well-developed κ-carbides are

    investigated. To ensure the compositional accuracy of APT measurements, the

    influence of measurement parameters on apparent compositions are systematically

    studied and the measurement conditions are optimized.

    Commonly, κ-carbides are regarded as coherent precipitates according to their

    nanometer-scale sizes observed by DF-TEM imaging and small lattice parameter

    difference with that of the -matrix based on XRD measurements (Sato et al., 1988;

    Choo et al., 1997; Bartlett et al., 2014). To my knowledge, there is no direct atomic-

    scale observation of the κ/ interface with respect to compositional and structural

    transitions. Due to the small lattice misfit between the κ-carbides and -matrix, the κ-

    peaks often appear close to the -peaks as small shoulders. Thus, an accurate

    determination of the lattice misfit between κ and is challenging. Moreover, to

    understand the non-uniform anisotropic κ/ microstructure (section 3.2.1), the interfaces

    between cuboidal/plate-like κ-carbides and broad/narrow -channels have to be

    individually studied. Possibly existing local incoherency cannot be detected via XRD.

    The existence of misfit dislocation at some point, however, could interact with mobile

    dislocations and affect their movement. Hence, another primary aim of this chapter is

    to gain a thorough view of the κ/ interfaces at the atomic scale. SXRD is used for an

    accurate evaluation of lattice misfit; APT and HAADF-STEM are exploited to study

    the interface structure.

    4.2 Results

    4.2.1 κ/ elemental partitioning and phase compositions

    The reconstructed 3D elemental maps for the SS-treated specimens are shown in

    Figure 4.1(a). All elements, i.e. Fe, Mn, Al and C, are homogeneously distributed

  • Chapter 4. Elemental partitioning and κ/ interface structure

    33

    throughout the specimen without any sign of clustering or phase separation. For a better

    assessment of the homogeneity of the elemental distributions, nearest neighbor (NN)

    distributions (Gault et al., 2012b) are evaluated (Figure 4.1(b)), where the

    experimentally measured distributions are compared against artificially randomized, i.e.

    correlation-free datasets. For the first as well as higher order NN distribution analyses,

    the experimental frequency histograms show a very good match with ideal random

    distributions for all detected elements. No elemental partitioning is detected for the SS-

    treated state. However, in the same SS-treated alloy, Welsch observed weak superlattice

    diffraction spots in TEM and finely dispersed ordered regions in a size of

  • Chapter 4. Elemental partitioning and κ/ interface structure

    34

    Table 4.1. Chemical compositions (at.%) of supersaturated single austenitic SS-treated

    specimens cnand -carbides cand -matrix c in the 24h-aged specimens determined

    by APT. Error bars indicate compositional fluctuations between different APT

    measurements, where the statistical errors are negligible.

    Fe Mn Al C

    cn 52.6±0.3 24.5±0.5 17.0±0.2 5.9±0.3

    c 43.4±0.3 23.7±0.3 19.8±0.2 13.2±0.4

    c 53.8±0.2 25.8±0.3 16.1±0.1 4.3±0.2

    After isothermal ageing at 600°C for 24 hours, κ-carbide precipitates are clearly

    observed both in TEM as ordered particles (Welsch, 2016) and in APT as a C-enriched

    phase (Figure 4.2(a), delineated by a C iso-concentration surface at a threshold value of

    9.0 at.%). Figure 4.2(b) shows a 1D concentration profile along a particle stack through

    a cylindrical ROI, which is highlighted in green in Figure 4.2(a). Al and C atoms clearly

    partition to the κ-carbides, whereas Mn exhibits only slight partitioning to . In order to

    closely compare the elemental partitioning of smaller plate-like and larger cuboidal κ-

    carbides with respect to their neighborhoods, proxigrams are employed to calculate the

    concentrations at fixed distances from their respective iso-concentration surfaces (9 at.%

    C). Figure 4.2(c) depicts proxigrams from one large cuboidal κ-carbide κ1L and three

    small plate-like κ-carbides κ1S, κ2

    S and κ3S, which are highlighted in blue in Figure

    4.2(a). The proxigrams of the plate-like particles show slightly higher statistical errors

    due to smaller probed volumes as compared to the large cuboidal κ-carbide, but no

    compositional difference related to precipitate size and shape is observed. All

    compositional profiles are rather smooth and symmetric. No elemental segregation or

    pile-up at κ/ interfaces is noticed.

    When evaluating phase compositions, contributions from overlapping mass-to-

    charge peaks must be considered. Those are however not taken into account in the

    above-shown concentration profiles and proxigrams. For a more accurate composition

    analysis, decomposition of overlapping mass-to-charge peaks has to be performed.

    Hence, phase compositions were determined by clipping out the phases in the 3D atom

    maps by iso-concentration surfaces and subsequently analyzing their individual mass

    spectra by decomposition of overlapping mass peaks (C2+/C4

    2+ and Al+/ Fe2+).

  • Chapter 4. Elemental partitioning and κ/ interface structure

    35

    Threshold values for iso-concentration surfaces that isolate the κ-carbides and -matrix

    were chosen as 10 at.% C and ≤7 at.% C, respectively. These values correspond to the

    respective plateau of the C concentrations in κ and in the 1D concentration profiles

    and exclude the interfacial region from the analyses (Figure 4.2(b)). Since no

    compositional differences were found between κ-carbides of different size and shape,

    an average composition value from a